Abstract
Alzheimer’s disease (AD) is the most common neurodegenerative disorder and there is currently no cure. Neural circuit dysfunction is the fundamental mechanism underlying the learning and memory deficits in patients with AD. Therefore, it is important to understand the structural features and mechanisms underlying the deregulated circuits during AD progression, by which new tools for intervention can be developed. Here, we briefly summarize the most recently established cutting-edge experimental approaches and key techniques that enable neural circuit tracing and manipulation of their activity. We also discuss the advantages and limitations of these approaches. Finally, we review the applications of these techniques in the discovery of circuit mechanisms underlying β-amyloid and tau pathologies during AD progression, and as well as the strategies for targeted AD treatments.
Keywords: Neural circuit, Alzheimer’s disease, Single cell RNA sequencing, Neural circuit tracing, Optogenetics, Chemogenetics
Introduction
Alzheimer’s disease (AD) is the most common neurodegenerative disorder in the elderly. The characteristic clinical symptom is progressive memory loss [1]. Several key neuropathological hallmarks of AD have been revealed, including extracellular neuritic plaques composed of β-amyloid (Aβ) and intracellular neurofibrillary tangles (NFTs) composed of hyperphosphorylated tau protein [2]. Both Aβ plaques and NFTs impair synaptic function and neural circuit networks. Using functional MRI, electrophysiological techniques, and biochemical data, researchers have found aberrant neural activities in patients with AD and mouse models [3, 4]. These abnormal activities trigger a disorganized brain network and interfere with the intricate processes underlying learning, memory, and other cognitive functions. Pharmacological therapies targeting either Aβ or tau have shown limited efficacy in the treatment of AD. This has motivated novel approaches designed not only to specifically rescue the cognitive impairments in AD but also avoid drug-induced side-effects or pharmacoresistant symptoms.
Different types of brain cells orchestrate to modulate cognitive function. Heterogeneous brain cells show distinct alterations during AD development. Given the advantages of neuron type-specificity and manipulation accuracy, single-cell RNA sequencing, optogenetics, and chemogenetics are increasingly used in the study of AD to investigate molecular and circuit mechanisms. Novel findings from AD studies that use these cutting-edge techniques may help us better understand the AD memory dysfunction induced by Aβ and tau proteins. They may also provide novel therapeutic strategies for AD treatment in the future. In this review, we briefly describe the most advanced tools currently available for neural circuit research relating to AD.
Single-Cell RNA Sequencing
There are billions of highly differentiated and interconnected cells within the brain. Distinct neurons communicate with each other and form intricate neural networks to control physical and mental functions. Traditional research at the bulk-tissue level of resolution may mask the complexity of changes across brain cells and within cell groups [5]. For example, using bulk RNA-sequencing, it may be difficult to differentiate whether low-abundance transcripts are expressed at low levels in common cell types or at high levels in rare cell types. Moreover, in diseases, the low abundance transcriptional changes in a rare cell type may be undetectable when a bulk RNA-sequencing approach is employed. As such, the newly developed single-cell RNA sequencing (scRNA-seq) techniques have the potential to overcome these limitations. In addition to providing a comprehensive landscape of brain cell-type diversity based on transcriptional profiles, the scRNA-seq method may be useful in mapping the mammalian connectome and exploring the mechanisms of brain-related disorders, such as AD, at the single-cell level.
Individual cells can be physically separated. The isolation methods vary in the number of isolated cells (high-throughput or low-throughput) and the means of cell selection (biased or unbiased) [6]. By whole-cell patch clamping, cellular content from a sucked single cell can be used to acquire RNA for quantitative RT-PCR or microarray analysis [7, 8]. Moreover, this method provides additional information, such as location, electrophysiological properties, and morphology of the examined cell. However, it is not quantitative enough, nor does it provide an unbiased assessment of global gene expression patterns. Moving from pipetting and manual selection, several automated single-cell compartmentalization methods have been developed, such as fluorescence-activated cell sorting and the fluidigm C1 system. Although they have been combined with scRNA-seq techniques, including SMART-seq and CEL-seq, to explore the brain at the single-cell level [9–11], there are two limitations: (1) the total number of captured cells is not quantitative enough for the high-throughput format; and (2) they do not maintain the native heterogeneity of brain constituents [12, 13]. Currently, droplet-based technologies to randomly capture single cells with barcoded beads are commonly used as a high-throughput and unbiased solution [14]. The three most widely used platforms are Droplet sequencing (Drop-seq) [15], indexing droplets RNA sequencing (inDrop) [16], and GemCode/Chromium 10× (widely known as 10× Genomics) [17, 18]. They use microfluidics to tag individual cells with single beads containing a unique barcode. Each mRNA transcript is also linked with a unique molecular identifier (UMI). By comparison, 10× Genomics is the most sensitive for detecting the greatest number of transcripts, while inDrop is ideal for detecting weakly-expressed genes (DEGs) [17].
scRNA-seq techniques require freshly-harvested, viable, and intact cells. Moreover, most of the single-cell isolation methods may impair neurons and cause some cell types to be under- or over-represented in the final datasets. Since nuclei are more resistant to mechanical and physical stress, single nuclei RNA sequencing (snRNA-seq) is an alternative approach to scRNA-seq and has been used in the study of the cellular composition in the brain [19–22]. Compared with whole-cell transcriptomes, there is no difference in snRNA-seq data in terms of the number of detected genes and resolved cell types [23].
Full-length and tag-based approaches are used to generate single-cell sequencing libraries [24]. Since reads are derived from across the entire length of genes, full-length methods greatly improve the overall sensitivity. It is worth noting that full-length library preparation shows a bias for longer genes rather than shorter genes because counts are often missed for the latter [25]. In tag-based methods, the incorporated UMIs allow the identification and quantification of individual transcripts.
A generalized single-cell transcriptomic experiment includes single cell isolation, barcoding, cDNA amplification, library construction, sequencing, and analysis. For each experiment, the technical performance of the scRNA-seq data should be evaluated from the sensitivity, accuracy and precision aspects [24, 26, 27]. Sensitivity measurement is highly dependent on the depth of sequencing [24, 26]. Accuracy can be influenced by factors specific to the protocol being used [24, 26]. Precision is inversely proportional to the technical noise in the RNA-seq measurements [24, 26].
A major disadvantage of the scRNA-seq method is the lack of spatial information [28–30]. By fluorescence in situ hybridization (FISH), the spatial locations of genes identified in scRNA-seq can be clearly visualized. For example, in single-molecule FISH (smFISH), multiple fluorescent probes are used to characterize distinct cell groups [31]. However, the number of colors that can be visualized at once is very limited in this technique. In contrast, sequential FISH (seqFISH) can sequentially label tissue for different RNA markers [32, 33]. After each round of labeling, fluorescence is washed out and cell populations are binarily labeled for the presence or absence of a particular gene. As such, seqFISH is more efficient and robust in cell population identification than smFISH.
Cutting-Edge Experimental Approaches for Visualizing Neural Circuits
To understand the complex brain functions carried out by neural circuits, it is necessary to dissect their structural organization. To date, several viral tracers have become available to delineate the structural connectivity between distinct neurons or different brain regions
Anterograde Trans-Synaptic Tracing
Herpes simplex virus (HSV) and vesicular stomatitis virus (VSV) have been commonly used in anterograde trans-synaptic tracing. Both viruses invade and replicate in first-order infected starter neurons and then spread into synaptically connected neurons.
HSV
The HSV type 1 strain H129 has been widely used in anterograde multi-synaptic tracing [34–36]. Tyrosine kinase (TK) is required for HSV replication and is necessary for its anterograde spread in non-mitotic cells. Recently, HSV has been genetically modified to achieve monosynaptic tracing. In this system, TK is deleted and then trans-complementation of TK restores the ability of H129-ΔTK to replicate and spread into directly projected targets. If the expression cassette of the TK gene is in a Cre-recombinase-dependent strategy, the starter cell type can be greatly restricted.
H129 mainly transmits anterogradely, but it can also infect upstream neurons through axon terminal uptake [37]. More importantly, the main drawback of the H129 strain is its high toxicity. It can also cause cell death and animal death. With current trans-multisynaptic H129 tracer systems, experimental animals die within 3 to 5 days after injection of H129 [38], which greatly restricts the sampling time to within 3 days. Although toxicity can be reduced by engineering viral genomes, such as TK-null H129 [39], this problem has not been resolved completely. TK-deficient recombinant viruses show milder cytotoxicity without damage to their infectivity and viral gene expression. However, trans-monosynaptic H129 tracer systems remain insufficient for electrophysiological recordings from brain slices. Thus, to support functional studies, more efforts are needed to reduce the cytotoxicity of trans-synaptic HSV in the future.
VSV
VSV is an arthropod-borne virus. It leads to vesiculation and ulceration around the mouth, teats, and hoofs after infection. These symptoms usually resolve within a few weeks without fatality [40]. Compared to HSV, VSV is less virulent to humans [41].
The glycoprotein (VSV-G) is necessary for VSV binding to target cells [42]. When VSV-G binds to phosphatidylserine on the cell surface membrane [41], VSV can enter cells and start to replicate rapidly. The first progeny of viruses are generated and released within 1.5 h [43]. It has been found that the direction of VSV transmission greatly depends on the glycoprotein. The intrinsic glycoprotein endows its characteristic of anterograde transmission [44]. However, G protein from either lymphocytic choriomeningitis virus or Rabies virus can change the transmission from anterograde to retrograde [44]. Toxicity and inefficiency are the main shortcomings of VSVs. Thus, the modification of virulence and efficiency of spread will enable the wider applications of VSV in circuit structure dissection and functional studies.
A recent study showed that adenovirus associated virus 1 (AAV1) and AAV9 can be used as monosynaptic anterograde tracers [45]. They exhibit anterograde trans-synaptic spread properties with relatively low efficiency. Therefore, the viral spread is restricted to direct postsynaptic targets of the infected cells, and the labelling of postsynaptic cells relies on reporter expression activated by a recombinase-expressing virus. Such a strategy can be used to perform afferent-dependent tagging of postsynaptic neurons and support functional manipulations with Cre-dependent tool genes. However, care should be taken when using it to map connections between reciprocally connected regions because retrograde trans-synaptic spread can also occur, especially at high titers [46].
Retrograde Trans-Synaptic Tracing
Rabies Virus (RV) is a neurotropic virus. It retrogradely transmits from the infected peripheral site to the central nervous system, leading to lethal zoonotic disease. Unlike HSV, the RV spreads exclusively in the retrograde direction. High neuronal specificity without causing glial cell infection makes RV an ideal tool for neural circuit tracing.
RV envelope protein G is necessary for neuronal infection and trans-synaptic transmission of the RV. SAD-B19, an attenuated vaccine strain of RV, was developed for the monosynaptic trans-synaptic system [47, 48]. In this system, the RVG gene is replaced by EGFP and pseudotyped with envelope glycoprotein (EnvA) from avian sarcoma-leukosis virus (ASLV-A). TVA is the cognate receptor of EnvA and does not exist in mammals. Thus, the recombinant RV [SADDG-EGFP(EnvA)] cannot infect mammalian neurons unless they are exogenously supplied with TVA. RV-DG(EnvA), a mutant rabies virus with glycoprotein G gene deletion and a helper virus (AAV) that expresses TVA and RVG in a Cre-dependent manner are both injected into the desired region, where RV can only invade cells with Cre expression and then retrogradely spread to their direct upstream neurons, but cannot spread any further. Therefore, in such cases, only monosynaptic connections to starter cells are marked [49].
Recently, Schwarz et al. developed a method termed TRIO (for tracing the relationship between input and output) and cell-type-specific TRIO (cTRIO) [50]. In this system, canine adenovirus type 2 (CAV2) is used to provide the recombinase (Cre or Flp). Although CAV2 is not a trans-synaptic tracer, it can transduce neurons through their axonal terminals and efficiently transport back to the soma. In the TRIO system, after the injection of CAV-Cre into one of the output regions, Cre is specifically expressed in neurons that project to the CAV2-injected region. Then, by injecting RVDG(EnvA) and Cre-dependent AAV helper virus that expresses TVA and RVG into the targeted area, RV can only infect neurons in the targeted area which directly project to the CAV2-injected regions. It then transports to monosynaptic upstream neurons. In contrast to TRIO, cTRIO has the advantage of being neuron type-specific. CAV2 is used to express Flp and AAV is used to express TVA. RVG is only in neurons with both Cre and Flp. With the help of these two viruses, RV can infect Cre-expressing neurons in the targeted area that innervate neurons within the CAV-injected region (Fig. 1). Thus, cTRIO enables monosynaptic tracing from a specific type of cell projecting to a specific brain area.
Cytotoxicity is the main limitation of the above RV tracing system (first generation). Chatterjee et al. recently introduced a new class of double-deletion-mutant rabies viral vectors, which show low toxicity in transduced cells [51]. They deleted a second gene known as ‘large protein’ gene (NCBI symbol: RABVgp5). This gene encodes the viral polymerase, which is required for the transcription of viral genes and replication of the viral genome [52]. Compared to first-generation viruses, these new (‘second-generation’ or ‘Δ GL’) viral vectors can also retrogradely infect projection neurons. The target neurons remain alive and healthy in terms of both electrophysiology and morphology up to a year after the virus injection. Thus, this system is promising for long-term tracing, functional detection, and optogenetic manipulation.
Cutting-Edge Experimental Approaches for Manipulating Neural Circuits
Optogenetics
Optogenetics could be an invaluable tool in neuroscience research. It encompasses the knowledge of optics, microbial biology, virology, and biochemistry [53, 54]. About 50 years ago, light-activated proteins were first discovered. Bacteriorhodopsin was discovered in 1971 [54], halorhodopsin in 1977 [55], and channelrhodopsin in 2002 [56]. Upon light exposure, these opsins act as channels or pumps to directly induce electrochemical signaling in cells. This feature is distinct from rhodopsin, which indirectly transduces electrical current via G-proteins. In 2005, a microbial opsin gene was first introduced into neurons and it precisely controlled neuronal spiking upon light exposure [57]. In 2007, the fiber-optic interface and single-component control of freely-moving mammals were described [58, 59]. A series of Cre recombinase-dependent opsin-expressing viruses have been developed and widely used in mouse lines that selectively express Cre recombinase in defined cell types. Thus, researchers can use these latest optogenetic controls to investigate the functions of defined cells in specific brain regions under physiological conditions and their alterations in diseases.
Channelrhodopsins, derived from Chlamydomonas reinhardtii, are one kind of light-driven proteins. Channelrhodopsin 2 (ChR2) is composed of seven transmembrane helices, and a retinal component is embedded within the helices [60]. Upon illumination, the retinal component absorbs light and quickly causes a conformational change, which opens the pore and allows the movement of cations down their electrochemical gradient on ChR2-expressing neurons. These inward currents drive fast membrane depolarization and robustly excite neurons [61, 62] (Fig. 2A). H134R, a gain-of-function mutant of ChR2, causes larger stationary photocurrents than wild-type ChR2 [63]. However, the recovery from inactivation of both wild-type ChR2 and H134R is slow, which limits the precision in the optogenetic control of neurons, especially at stimulating frequencies above the γ range [57, 64]. Gunaydin et al. designed point mutations in the ChR2 sequence and developed ChETA, which accelerates channel closing and rapid repolarization of neurons following light stimulation [65]. Recently, many variants have been developed, such as red-shifted (VChR1 [66], ReaChR [67]) and blue-shifted (CheRiff) variants [68].
In addition to depolarizing cells, some microbial opsins can hyperpolarize cells upon illumination. NpHR isolated from Natronomonas pharaonis is sensitive to yellow-green light. Light activation of NpHR produces an inward Cl– (hyperpolarizing current), which inhibits the action potential firing of NpHR-expressing neurons (Fig. 2B). Arch is a light-activated proton pump bacteriorhodopsin [69]. By pumping H+ out, Arch also hyperpolarizes cells when exposed to light (Fig. 2B). Furthermore, ChR2 can also be used to inhibit neurons upon light stimulation when its permeability is changed to Cl– instead of cations after mutation in its pore structure [70] (Fig. 2B).
Chemogenetics
Recently, chemogenetics has been widely employed in neuroscience research to regulate the activities of specific neurons or neural circuits. Chemogenetics refers to the use of genetically-engineered receptors that interact with specific synthetic ligands or molecules to alter cellular signal transduction [71]. Since the receptors (e.g. G protein-coupled receptors and ligand-gated ion channels) are modified through random or site-directed mutagenesis, they are no longer responsive to their natural ligands, but can be specifically activated by synthetic chemicals [71].
Designer receptors exclusively activated by designer drugs (DREADDs) [72] are engineered G protein-coupled receptors (GPCRs) that can precisely control GPCR signaling pathways (for example, Gq, Gs, and Gi). Currently, hM3Dq and hM4Di are the most commonly-used DREADDs [73, 74]. Due to their mutations, both hM3Dq and hM4Di can only be activated in the presence of clozapine N-oxide (CNO), a compound modified from clozapine (Fig. 2C, D). CNO only activates these “designer receptors” without changing the functions of endogenous receptors. However, this specificity can be changed at a high dose of CNO. It has been reported that intracranial injections of CNO at micromolar concentrations (10 μmol/L) competitively inhibit binding at several receptors, including muscarinic M1, M3, M4, histamine H1, 5-HT2A, and dopamine D1 and D2 [75]. Thus, these off-target effects at endogenous receptors may significantly confound the agonistic effects of CNO.
Compared to optogenetics, chemogenetics has an incomparable advantage in that the stimulus can be administered via less- or non-invasive routes, such as intraperitoneal injection and oral administration [73, 76, 77]. Although chemogenetics cannot achieve precise temporal control, they can provide prolonged manipulation of neurons and their circuit activities [77].
Neural Circuit Defect in AD Brains and Strategies for Targeted Manipulations
The hippocampus is vulnerable in patients with AD [78–80], and its dysfunction is closely associated with cognitive impairment during AD progression [80, 81]. Next, we focus on the hippocampus and briefly introduce some recent studies on hippocampus-associated neural circuits in AD.
Application of Single Cell Transcriptomics in AD
In addition to cell-type classification under physiological conditions, snRNA-seq has recently been used in AD to understand the vulnerability of different cell types [19, 82]. 80,660 droplet-based single-nucleus transcriptomes from prefrontal cortex in human subjects with varying degrees of AD pathology were profiled and analyzed [19]. Differentially-expressed genes between AD-pathology and no-pathology groups were downregulated in excitatory (Ex) and inhibitory (In) neurons, while they were upregulated in oligodendrocytes (Olis), astrocytes (Asts), and microglia (Mic), indicating a heterogeneous response to AD pathologies among cell types [19]. In cellular subpopulations, many subclusters of Ex and In neurons, Asts, Mic, Olis, and oligodendrocyte progenitor cells (Opcs), have been identified. Among them, some cellular subpopulations were associated with cells isolated from subjects with AD pathology traits, while some were associated with cells from subjects with no pathology [19]. Furthermore, genes associated with protein folding and stability, neuronal and necrotic death, T-cell activation, and immunity were identified in some subpopulations of AD pathology-associated In neurons, Opcs, and Mic, indicating cell-type-specific responses in AD [19]. Moreover, male and female individuals showed differential transcriptional responses to AD pathology, especially in neurons and oligodendrocyte cells [19]. Thus, this single-cell view of transcriptional alterations associated with AD pathologies highlights the complexity of AD mechanisms from the perspective of cellular heterogeneity. Also, an understanding of neuronal vulnerability may expedite the study of circuit mechanisms in AD and provide molecular targets for circuit-specific interventions.
Glutamatergic Circuits
Dysfunction of Glutamatergic Neurons in the Hippocampus During AD
In AD, Aβ can induce neuronal hyperactivation and hypoactivation. Aβ-dependent neuronal hyperactivation may contribute to preclinical hippocampal hyperexcitability [83–85]. Mechanistically, Aβ inhibits the glutamate uptake capacity of astrocytes to produce excess glutamate accumulation, leading to the neurotoxic effects in AD [86, 87]. On the other hand, oligomeric Aβ acts on α-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid receptors (AMPARs) [88] and extrasynaptic N-methyl-D-aspartate receptors (NMDARs) to elicit excitotoxic effects [89, 90]. In turn, neuronal hyperactivity promotes the activity-dependent production of Aβ, eventually forming a positive feedback and leading to synapse loss, arbor shrinkage, and hypoexcitability in the late stages of AD [91–93]. In postmortem AD brains, Aβ is co-localized in glutamatergic boutons. Aβ accumulation causes loss of dendritic spines and severe neuropil damage [94–97]. Functionally, excess Aβ gains toxicity by inhibiting synaptic transmission [94, 98, 99], and impairing synaptic plasticity [94, 96, 100–103].
Tau pathology is another causal factor leading to the impairment of glutamatergic synapses. In our previous study, we found that overexpression of wild-type human tau (htau) in hippocampal CA3 decreases spontaneous excitatory postsynaptic currents with unchanged spontaneous inhibitory postsynaptic currents. These results indicate that glutamatergic synapses in hippocampal CA3 are more vulnerable to tau pathology [104]. Mechanistically, calcineurin-mediated inactivation of nuclear CaMKIV/CREB signaling is responsible for the glutamatergic impairments and memory deficits induced by htau accumulation[104]. When tau accumulates in the hippocampus, it inhibits NMDAR expression by upregulating STAT1, a transcription factor [105]. It also inhibits CREB/GluN1 phosphorylation by suppressing PKA [106]. In addition to accumulation, tau is mis-localized from the axon to the dendrites during AD progression. Then, the mis-localized tau interacts with fyn and triggers fyn to move to the dendritic spines to phosphorylate the GluN2B subunit of NMDAR, thereby enhancing excitotoxicity [107]. In AD, tau undergoes aberrant post-translational modifications, such as hyperphosphorylation, sumoylation, acetylation, glycosylation, andproteolytic cleavages, which promote tau accumulation and hippocampal propagation, causing memory deficits [108–111]. In cultured hippocampal neurons, expression of mutated tau, which mimics the phosphorylation of 14 residues, significantly reduces the number of AMPARs at synapses [112]. When expressing htau with mutations to mimic the acetylation of K274 and K281, mice have deficits in spatial and pattern separation memory with impairment in AMPAR trafficking during LTP [113]. Blocking the caspase-2 cleavage of human P301L tau by mutating the Asp314 residue on tau restores AMPAR-mediated synaptic transmission in neurons and memory deficits in mice [114]. Although P301L tau has been found to reduce the excitability of hippocampal CA1 neurons by shifting the axon initial segment (AIS) [115], Sohn et al. found that the V337M tau mutation in human neurons leads to hyperexcitability of neuronal networks by impacting the AIS cytoskeleton [116].
Recently, an in vivo neural dynamic recording system has been developed to detect cell-specific neuronal activity while monitoring the behaviors of freely-moving subjects. For example, calcium-based fiber photometry is one such system that measures calcium fluorescence signals and directly reflects neuronal spiking activity in vivo. In APP/PS1 mice, a genetically encoded calcium indicator, GCaMP6 [117], was expressed in basolateral amygdala (BLA) neurons. During the elevated plus maze test, disorganized firing patterns of BLA neurons were clearly detectable in response to BLA Aβ pathology and correlated with the anxious state of APP/PS1 mice [34]. Thus, in addition to hyperactivation and hypoactivation, AD pathologies produce abnormal neuronal firing patterns, which may result in dysfunction of information encoding and contribute to AD phenotypes.
It is worth noting that the fiber photometry system can only record population calcium. A fluorescence microscope (miniscope) system has recently been applied in neuroscience research to visualize neural activity from deep brain regions in freely-behaving animals [118]. The advantage of this calcium imaging is that the recording of neural activity from a specific population of neurons can occur longitudinally at the single-cell level of resolution. It is expected that the combination of a fluorescence microscope system and behavioral tests will provide more detailed information on neuronal dysfunction and cognitive impairment in AD.
Targeting Glutamatergic Neurons to Rescue AD-Like Memory Disorder
Whether and how targeting abnormal glutamatergic neurons could restore memory deficits remain unclear. Previous studies have suggested that episodic memory deficits in patients with AD may be due to information encoding failure [119]. However, Roy et al. employed optogenetics combined with behavioral testing and found significant disruption of memory retrieval in the early stages of AD [120]. They used immediate-early gene-dependent tools [121] to label memory engram cells (i.e., neurons holding traces of a specific memory) in APP/PS1 mice (an AD mouse model) and reported a close correlation between the progressive reduction in spine density of the hippocampal dentate gyrus (DG) engram cells and age-dependent amnesia [120]. Direct photoactivation of hippocampal DG memory engram cells with ChR2 was sufficient to induce memory recall in AD mice, indicating a deficit of memory retrieval during early AD-related memory loss [120]. Further, optogenetic induction of LTP at perforant path synapses of hippocampal DG engram cells significantly restored both spine density and memory in AD mice [120]. This study highlights the contribution of memory retrieval failure during early AD-related memory loss. Moreover, it identifies a causal role of entorhinal cortex-DG glutamatergic circuit dysfunction in memory retrieval disruption in AD.
Glutamatergic neurons in the hippocampus also receive innervation from the amygdala and their connections undergo robust degeneration during AD. Under physiological conditions, researchers combined retrograde (RV [122]) and anterograde (HSV [34]) tracing systems to delineate a novel connection between the posterior BLA (pBLA) and ventral hippocampal CA1 (vCA1), that is, the pBLA-vCA1 circuit. Using optogenetics and electrophysiological recordings, they demonstrated that vCA1 Calbindin 1 (vCA1Calb1) neurons are downstream of pBLA glutamatergic neurons [34]. However, in AD mice, the pBLA-vCA1Calb1 circuit is significantly inhibited in response to Aβ accumulation and results in AD-like anxiety. Photostimulation of the pBLA-vCA1Calb1 circuit efficiently attenuates AD-like anxiety. Simultaneously, targeting the pBLA-vCA1Calb1 circuit by optogenetics during memory retrieval significantly restores memory disorders in AD mice [34]. This novel circuit links emotions to memory. It may also provide a promising intervention target for AD patients with anxiety and cognitive impairment.
Previous work has revealed that elevated neuronal activity accelerates the progression of pathological Aβ and tau [123] and in turn continuously destroys circuit functions, forming a vicious cycle. In EC-Tau/hAPP mice, hM4Di was virally delivered and expressed in the EC, then, CNO was intraperitoneally injected. Using these chemogenetic tools, neuronal activity in the entorhinal cortex was significantly attenuated, and its beneficial effects on reducing Aβ accumulation and tau spreading into the hippocampus along the cortex-hippocampal (EC-HP) circuit were detected [77]. Thus, inhibition of hyperactivated circuits in response to Aβ and tau may be a novel strategy to prevent the progression of AD along the neural networks and break the vicious cycle between AD pathologies and memory-associated circuit dysfunction.
GABAergic Circuits
Dysfunction of GABAergic Neurons in the Hippocampus During AD
The GABA is the main inhibitory neurotransmitter in the brain. Early studies in postmortem human brains and animal models revealed that GABAergic neurons are less vulnerable than glutamatergic neurons to AD pathology. However, this idea has been challenged. There is accumulating evidence that GABAergic neurotransmission undergoes enormous changes in AD, causing excitatory/inhibitory (E/I) imbalance during AD progression.
In brain sections from patients with AD and APP/PS1 mice, perisomatic GABAergic terminals are significantly decreased, especially on the cortical neurons adjacent to amyloid plaques, suggesting an association between GABAergic neuronal impairments and Aβ [124, 125]. In addition to loss, GABAergic presynaptic terminals are elevated at an early stage in tgCRND8 and APP/PS1 mice [126, 127], indicating complicated alterations of GABAergic function in response to Aβ pathologies. During AD progression, GABAergic neuronal subpopulations are also lost. Both somatostatin ‐expressing neurons (conferring distal dendritic inhibition of pyramidal cells) and parvalbumin (PV)‐expressing neurons (providing perisomatic inhibition of pyramidal cells) are notably decreased in the brains of patients with AD and transgenic mice with Aβ pathologies [128–130].
There is a reciprocity between tau hyperphosphorylation and GABAergic synaptic dysfunction. On one hand, GABAergic signaling regulates tau hyperphosphorylation. Several GABAA receptor modulators have been found to increase the interaction between tau and Peptidyl-prolyl cis-trans isomerase 1 (Pin 1), by which more protein phosphatase 2A (PP2A) is recruited to the cell surface to dephosphorylate GABAA receptor β3 subunit and simultaneously reduces the availability of PP2A for tau dephosphorylation [131]. In line with this mechanism, the activation of GABAA receptors robustly induces tau hyperphosphorylation at the AT8 epitope in cultured cortical neurons [131]. On the other hand, tau pathologies modify GABA release. In P301L mice, GABAergic interneurons are hyperactivated, resulting in higher GABA levels in the brain [132]. Thus, there may be a vicious cycle by which activation of GABAA receptors promote tau hyperphosphorylation by reducing the association of PP2A with tau, then, the hyperphosphorylated tau enhances GABAergic neurotransmission in turn during AD progression.
Dysfunction of GABAergic transmission produces E/I imbalance in local circuits. In 14-month-old APdE9 mice, hyperexcitability was detected in the hippocampal DG and this was associated with the silencing of local inhibitory neurons [133]. Recently, we reported a prominent accumulation of hyperphosphorylated tau in a DG subset in AD patients and mice, including mature excitatory neurons, immature granular neurons, and GABAergic interneurons. However, only interneuron-specific overexpression of full-length wild-type htau to mimic AD-like tau accumulation in the mouse DG induced adult hippocampal neurogenesis deficits and increased neural stem cell-derived astrogliosis. Using calcium-based fiber photometry, hyperactivated neighboring excitatory neurons were detected in vivo after interneuron-specific overexpression of htau. Chemogenetic inhibition of excitatory neurons or pharmacologically strengthening GABAergic efficiently rescue the human tau-induced deficits in adult hippocampal neurogenesis with improved contextual cognition [134]. This work demonstrated a causal role of tau accumulation in GABAergic impairments and linked local circuit disinhibition with disruption of hippocampal neurogenesis in AD.
In addition to local microcircuits, GABAergic neurons in CA1 are also innervated by pyramidal neurons in the entorhinal cortex layer II (ECIIPN) [135]. Using a retrograde monosynaptic tracing system, researchers have identified direct connections between ECIIPN and parvalbumin-positive CA1 (CA1PV, one type of GABAergic neuron) under physiological conditions. Their functional connection (ECIIPN → CA1PV) was significantly decreased according to optogenetic and electrophysiological examinations. Further, in vivo electrophysiological recording revealed a disruption of E/I balance and dispersed place-associated firings in CA1 during ECIIPN → CA1PV degeneration in AD mice. No change in ECIIPN → CA1PN (pyramidal neurons in CA1) was observed. These data highlight the vulnerability of the ECIIPN → CA1PV circuit and its role in encoding the impairment of place information in AD.
In addition, there are some long-range inhibitory circuits from other brain regions, such as the amygdala [136] and medial septum [137], to the hippocampus. However, the details of these connections, such as the neuron type, the innervation pattern, and their transcriptome characteristics, remain elusive. More importantly, whether and how their structure and function change during AD progression also deserve further investigation.
Targeting GABAergic Neurons to Rescue AD-Like Memory Disorders
Dysfunctions in GABAergic transmission can lead to theta and gamma oscillation impairments which are important for synaptic plasticity and spatial memory. Theta oscillations (6–10 Hz) have been linked to exploratory behavior-related activities, reflecting internally-generated dynamics [138] and goal-related behaviors [139]. In rTg4510 mice (a model of tauopathy), theta and place cell (i.e. cells firing at a specific location in an environment) sequences are both significantly impaired. Principal cell firing is independent of the environmental context,, indicating the inability of mice to form new spatial memories. A reduced firing rate of the recorded interneurons may contribute to this rigid firing sequence [140]. GABAergic cell loss can decrease theta power, which is associated with cognitive dysfunction in a rat amyloid model [141]. Impaired gamma oscillation integrity has also been reported in the AD model [142, 143] and patients [144], contributing to the inefficient execution of spatial working memory [145].
However, there are distinct subtypes of interneurons in the hippocampus. Whether and how targeting different hippocampal interneurons rescues the abnormal oscillations in AD and restores memory loss remains unclear. Somatostatin-positive (SST) interneurons preferentially modulate theta oscillations [146], while parvalbumin-positive (PV) interneurons modulate gamma oscillations [147, 148]. A reduction in behaviorally-driven gamma oscillations is detectable before the onset of plaque formation or cognitive decline in a mouse model of AD. Using a Cre-dependent strategy, optoactivation of PV interneurons at gamma (40 Hz) but not at other frequencies significantly reduces Aβ levels [147]. In the J20 AD mouse model, optostimulation of PV neurons in the medial septum specifically at 40 Hz during memory retrieval restores the hippocampal slow gamma oscillation (30-60 Hz) [149]. In Aβ-injected SST-Cre or PV-Cre mice, optogenetic activation of ChR2-expressing SST and PV interneurons increases theta power and gamma oscillations, respectively, with resynchronized CA1 pyramidal cell (PC) spikes. Photoactivation of SST and PV interneurons resynchronizes SST and PV interneuron spike phases relative to theta and gamma oscillations [150], while simultaneously selectively enhancing spontaneous inhibitory postsynaptic currents onto CA1 PCs at theta and gamma frequencies, respectively [150]. Given the degeneration of the ECIIPN–CA1PV pathway in AD, optogenetic activation of ECIIPN–CA1PV synapses with a theta burst stimulation paradigm once per day for 35 consecutive days effectively enhanced theta oscillations in the hippocampal CA1 of AD mice and interrupted the progression of ECIIPN–CA1PV synaptic decays [135]. Targeting the ECIIPN–CA1PV pathway also improved the representation and precise spatial firing of CA1 place cells in AD mice, simultaneously rescuing their spatial memory deficits [135]. Together, targeting distinct interneurons may be a potential therapeutic strategy for restoring disorganized hippocampal networks and synaptic plasticity impairments in AD.
Others
The hippocampus also receives cholinergic innervation [151]. Choline acetyltransferase (ChAT) is responsible for acetylcholine synthesis and is viewed as a reliable marker of cholinergic integrity [151]. In AD brains, cortical ChAT activity is significantly decreased, and this is closely correlated with the severity of dementia [152]. In the AD hippocampus, ChAT immunopositivity is also robustly decreased [153]. Thus, degeneration of cholinergic neurons in the basal forebrain may lead to a reduction in cholinergic input to the AD hippocampus. α-Secretase-cleaved APP has been found to accumulate in cholinergic dystrophic neurites during AD progression [154]. Overexpression of Aβ decreases cholinergic spontaneous and miniature excitatory postsynaptic currents [155]. When anti-murine-p75-SAP (conjugation of saporin to a rat monoclonal antibody against the mouse p75 nerve growth factor receptor) was intracerebroventricularly injected, ChAT activity in the hippocampus and neocortex was selectively inhibited in a dose-dependent manner. These anti-murine-p75-SAP-lesioned mice showed significant spatial memory deficits in the Morris water maze test therapies remain critical in the management of patients with AD [156], the precise structural connections between cholinergic neurons in the basal forebrain and hippocampal neurons, as well as their alterations during AD progression, remain unclear.
In AD, rates of cognitive deficits are also correlated with impairment of the serotonergic system [157, 158]. Serotonergic cells are lost in the brainstem of patients with AD [159]. Selective serotonin reuptake inhibitors improve cognitive function and reduce AD-associated behavioral disorders [159, 160]. Anatomically, serotonergic neurons consist mainly of dorsal and median raphe nuclei, which send dense serotonergic projections to the hippocampus [160, 161]. It has been found that serotonin inhibits the ventral hippocampus, and its inhibition is responsible for sustained goal-directed behavior [162]. Photostimulation of 5-HT neurons expressing ChR2 in the median raphe nucleus significantly increases 5-HT concentration in the dorsal hippocampus (dHC) and produce anxiety-like behavior [163]. Further, optoactivation of 5-HT terminals in the dHC also promote anxiety [163]. Within the hippocampus, optoactivation of serotonergic terminals in CA1 enhances excitatory transmission at CA3-to-CA1 synapses and promotes spatial memory in ChR2-expressing mice [164]. Optogenetic inhibition of CA1 5-HT terminals via Arch inhibits spatial memory [164]. However, how changes occur in the hippocampus-associated serotonergic circuits in AD remain elusive.
Perspective
The study of neural circuitry is of great significance for understanding the mechanisms linking brain pathologies and cognitive symptoms in AD. Here, we have summarized the major experimental approaches for neural circuit tracing and their functional manipulations. Furthermore, the introduction of aberrant neural circuits identified in AD mouse models or patients may provide novel therapeutic strategies for AD treatment in the clinic (Fig. 3). However, some key questions need to be addressed.
The application of viral approaches greatly facilitates research on AD neural circuit dissection. In addition to applications, we should pay more attention to their advantages and shortcomings, as noted earlier in this review. To date, there is no perfect viral vector that suits all needs. H129-ΔTK in anterograde trans-synaptic tracing cannot be used to trace local outputs because non-TK-expressing cells can also be infected at the injection site. Due to its virulence, HSV is not suitable for long-term tracing. Retrograde viral tracers, RV, also require further optimization to reduce their toxicity and meet the needs for electrophysiological recording. Furthermore, the infection and tracing efficiency of viral tracers may vary in different brain regions and cell types. Thus, cross-validating tracing results among multiple tracers is highly recommended for the dissection of AD-associated neural circuits.
Optogenetics and chemogenetics have enormous advantages in the functional studies of neural circuits. Combined with electrophysiology and tracers, we can identify functional connections between distinct neurons and evaluate their functional alterations in AD models. Together with behavioral tests, we can find specific neural circuits involved in different stages of the memory process and identify which circuit is responsible for AD-like cognitive dysfunction in various AD animal models. Optogenetics and chemogenetics have not only helped us understand the pathogenesis of AD but have also opened the door to using these tools for providing treatment. However, there are at least four major obstacles to the application of optogenetics and chemogenetics in patients with AD: (1) autoimmune reactions in patients because of the direct introduction of foreign objects into the brain, (2) the methodology of opsin and DREADD expression in specifically targeted neurons without invasion, (3) the evaluation of opsin and DREADD expression levels in specifically targeted neurons, and (4) tissue damage due to overheating from light stimulation.
Due to space constraints in this review, we introduced only some of the recent studies on hippocampal glutamatergic and GABAergic neural circuits in AD. Other glutamatergic and GABAergic neural circuits outside the hippocampal regions and neural circuits of other neuron types, such as cholinergic and serotonergic neurons, also deserve further investigation.
Acknowledgements
This review was supported by Grants from the Natural Science Foundation of China (31730035, 82071219, 91632305, and 91949205), the Ministry of Science and Technology of China (2016YFC1305800), and the Guangdong Provincial Key S&T Program (2018B030336001).
Contributor Information
Yang Ying, Email: yingyang@hust.edu.cn.
Jian-Zhi Wang, Email: wangjz@mail.hust.edu.cn.
References
- 1.Qiu C, Kivipelto M, von Strauss E. Epidemiology of Alzheimer's disease: Occurrence, determinants, and strategies toward intervention. Dialogues Clin Neurosci. 2009;11:111–128. doi: 10.31887/DCNS.2009.11.2/cqiu. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 2.Kumar A, Singh A, Ekavali. A review on Alzheimer's disease pathophysiology and its management: An update. Pharmacol Rep 2015, 67: 195–203. [DOI] [PubMed]
- 3.Gu XM, Jiang ZF, Huang HC. Magnetic resonance imaging of Alzheimer's disease: From diagnosis to therapeutic evaluation. Chin J Integr Med. 2010;16:276–282. doi: 10.1007/s11655-010-0276-8. [DOI] [PubMed] [Google Scholar]
- 4.Harrison TM, Maass A, Adams JN, Du R, Baker SL, Jagust WJ. Tau deposition is associated with functional isolation of the Hippocampus in aging. Nat Commun. 2019;10:4900. doi: 10.1038/s41467-019-12921-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 5.De Strooper B, Karran E. The cellular phase of Alzheimer's disease. Cell. 2016;164:603–615. doi: 10.1016/j.cell.2015.12.056. [DOI] [PubMed] [Google Scholar]
- 6.Svensson V, Vento-Tormo R, Teichmann SA. Exponential scaling of single-cell RNA-seq in the past decade. Nat Protoc. 2018;13:599–604. doi: 10.1038/nprot.2017.149. [DOI] [PubMed] [Google Scholar]
- 7.Rossier J, Bernard A, Cabungcal JH, Perrenoud Q, Savoye A, Gallopin T, et al. Cortical fast-spiking parvalbumin interneurons enwrapped in the perineuronal net express the metallopeptidases Adamts8, Adamts15 and Neprilysin. Mol Psychiatry. 2015;20:154–161. doi: 10.1038/mp.2014.162. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 8.Subkhankulova T, Yano K, Robinson HP, Livesey FJ. Grouping and classifying electrophysiologically-defined classes of neocortical neurons by single cell, whole-genome expression profiling. Front Mol Neurosci. 2010;3:10. doi: 10.3389/fnmol.2010.00010. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 9.Hashimshony T, Senderovich N, Avital G, Klochendler A, de Leeuw Y, Anavy L, et al. CEL-Seq2: Sensitive highly-multiplexed single-cell RNA-Seq. Genome Biol. 2016;17:77. doi: 10.1186/s13059-016-0938-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 10.Zeisel A, Muñoz-Manchado AB, Codeluppi S, Lönnerberg P, La Manno G, Juréus A, et al. Brain structure Cell types in the mouse cortex and Hippocampus revealed by single-cell RNA-seq. Science. 2015;347:1138–1142. doi: 10.1126/science.aaa1934. [DOI] [PubMed] [Google Scholar]
- 11.Usoskin D, Furlan A, Islam S, Abdo H, Lönnerberg P, Lou DH, et al. Unbiased classification of sensory neuron types by large-scale single-cell RNA sequencing. Nat Neurosci. 2015;18:145–153. doi: 10.1038/nn.3881. [DOI] [PubMed] [Google Scholar]
- 12.Wu AR, Neff NF, Kalisky T, Dalerba P, Treutlein B, Rothenberg ME, et al. Quantitative assessment of single-cell RNA-sequencing methods. Nat Methods. 2014;11:41–46. doi: 10.1038/nmeth.2694. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 13.Pollen AA, Nowakowski TJ, Shuga J, Wang XH, Leyrat AA, Lui JH, et al. Low-coverage single-cell mRNA sequencing reveals cellular heterogeneity and activated signaling pathways in developing cerebral cortex. Nat Biotechnol. 2014;32:1053–1058. doi: 10.1038/nbt.2967. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 14.Prakadan SM, Shalek AK, Weitz DA. Scaling by shrinking: Empowering single-cell ‘omics’ with microfluidic devices. Nat Rev Genet. 2017;18:345–361. doi: 10.1038/nrg.2017.15. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 15.Macosko EZ, Basu A, Satija R, Nemesh J, Shekhar K, Goldman M, et al. Highly parallel genome-wide expression profiling of individual cells using nanoliter droplets. Cell. 2015;161:1202–1214. doi: 10.1016/j.cell.2015.05.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 16.Klein AM, Mazutis L, Akartuna I, Tallapragada N, Veres A, Li V, et al. Droplet barcoding for single-cell transcriptomics applied to embryonic stem cells. Cell. 2015;161:1187–1201. doi: 10.1016/j.cell.2015.04.044. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 17.Zhang XN, Li TQ, Liu F, Chen YQ, Yao JC, Li ZY, et al. Comparative analysis of droplet-based ultra-high-throughput single-cell RNA-seq systems. Mol Cell. 2019;73:130–142.e5. doi: 10.1016/j.molcel.2018.10.020. [DOI] [PubMed] [Google Scholar]
- 18.Zheng GX, Terry JM, Belgrader P, Ryvkin P, Bent ZW, Wilson R, et al. Massively parallel digital transcriptional profiling of single cells. Nat Commun. 2017;8:14049. doi: 10.1038/ncomms14049. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 19.Mathys H, Davila-Velderrain J, Peng ZY, Gao F, Mohammadi S, Young JZ, et al. Author Correction: Single-cell transcriptomic analysis of Alzheimer's disease. Nature. 2019;571:E1. doi: 10.1038/s41586-019-1329-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 20.Lake BB, Ai R, Kaeser GE, Salathia NS, Yung YC, Liu R, et al. Neuronal subtypes and diversity revealed by single-nucleus RNA sequencing of the human brain. Science. 2016;352:1586–1590. doi: 10.1126/science.aaf1204. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 21.Krishnaswami SR, Grindberg RV, Novotny M, Venepally P, Lacar B, Bhutani K, et al. Using single nuclei for RNA-seq to capture the transcriptome of postmortem neurons. Nat Protoc. 2016;11:499–524. doi: 10.1038/nprot.2016.015. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 22.Lake BB, Chen S, Sos BC, Fan J, Kaeser GE, Yung YC, et al. Integrative single-cell analysis of transcriptional and epigenetic states in the human adult brain. Nat Biotechnol. 2018;36:70–80. doi: 10.1038/nbt.4038. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 23.Lake BB, Codeluppi S, Yung YC, Gao D, Chun J, Kharchenko PV, et al. A comparative strategy for single-nucleus and single-cell transcriptomes confirms accuracy in predicted cell-type expression from nuclear RNA. Sci Rep. 2017;7:6031. doi: 10.1038/s41598-017-04426-w. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 24.Ziegenhain C, Vieth B, Parekh S, Reinius B, Guillaumet-Adkins A, Smets M, et al. Comparative analysis of single-cell RNA sequencing methods. Mol Cell. 2017;65:631–643.e4. doi: 10.1016/j.molcel.2017.01.023. [DOI] [PubMed] [Google Scholar]
- 25.Phipson B, Zappia L, Oshlack A. Gene length and detection bias in single cell RNA sequencing protocols. F1000Res 2017, 6: 595. [DOI] [PMC free article] [PubMed]
- 26.Wu AR, Wang JB, Streets AM, Huang YY. Single-cell transcriptional analysis. Annu Rev Anal Chem (Palo Alto Calif) 2017;10:439–462. doi: 10.1146/annurev-anchem-061516-045228. [DOI] [PubMed] [Google Scholar]
- 27.Saliba AE, Westermann AJ, Gorski SA, Vogel J. Single-cell RNA-seq: Advances and future challenges. Nucleic Acids Res. 2014;42:8845–8860. doi: 10.1093/nar/gku555. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 28.Crosetto N, Bienko M, van Oudenaarden A. Spatially resolved transcriptomics and beyond. Nat Rev Genet. 2015;16:57–66. doi: 10.1038/nrg3832. [DOI] [PubMed] [Google Scholar]
- 29.Zeng HK, Sanes JR. Neuronal cell-type classification: Challenges, opportunities and the path forward. Nat Rev Neurosci. 2017;18:530–546. doi: 10.1038/nrn.2017.85. [DOI] [PubMed] [Google Scholar]
- 30.Lein E, Borm LE, Linnarsson S. The promise of spatial transcriptomics for neuroscience in the era of molecular cell typing. Science. 2017;358:64–69. doi: 10.1126/science.aan6827. [DOI] [PubMed] [Google Scholar]
- 31.Raj A, van den Bogaard P, Rifkin SA, van Oudenaarden A, Tyagi S. Imaging individual mRNA molecules using multiple singly labeled probes. Nat Methods. 2008;5:877–879. doi: 10.1038/nmeth.1253. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 32.Shah S, Lubeck E, Zhou W, Cai L. In situ transcription profiling of single cells reveals spatial organization of cells in the mouse Hippocampus. Neuron. 2016;92:342–357. doi: 10.1016/j.neuron.2016.10.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 33.Lubeck E, Coskun AF, Zhiyentayev T, Ahmad M, Cai L. Single-cell in situ RNA profiling by sequential hybridization. Nat Methods. 2014;11:360–361. doi: 10.1038/nmeth.2892. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 34.Pi GL, Gao D, Wu DQ, Wang YL, Lei HY, Zeng WB, et al. Posterior basolateral amygdala to ventral hippocampal CA1 drives approach behaviour to exert an anxiolytic effect. Nat Commun. 2020;11:183. doi: 10.1038/s41467-019-13919-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 35.Dum RP, Levinthal DJ, Strick PL. The spinothalamic system targets motor and sensory areas in the cerebral cortex of monkeys. J Neurosci. 2009;29:14223–14235. doi: 10.1523/JNEUROSCI.3398-09.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 36.McGovern AE, Davis-Poynter N, Farrell MJ, Mazzone SB. Transneuronal tracing of airways-related sensory circuitry using Herpes simplex virus 1, strain H129. Neuroscience. 2012;207:148–166. doi: 10.1016/j.neuroscience.2012.01.029. [DOI] [PubMed] [Google Scholar]
- 37.Su P, Wang HD, Xia JJ, Zhong X, Hu L, Li YL, et al. Evaluation of retrograde labeling profiles of HSV1 H129 anterograde tracer. J Chem Neuroanat. 2019;100:101662. doi: 10.1016/j.jchemneu.2019.101662. [DOI] [PubMed] [Google Scholar]
- 38.Wojaczynski GJ, Engel EA, Steren KE, Enquist LW, Patrick Card J. The neuroinvasive profiles of H129 (Herpes simplex virus type 1) recombinants with putative anterograde-only transneuronal spread properties. Brain Struct Funct. 2015;220:1395–1420. doi: 10.1007/s00429-014-0733-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 39.Kit S, Kit M, Pirtle EC. Attenuated properties of thymidine kinase-negative deletion mutant of pseudorabies virus. Am J Vet Res. 1985;46:1359–1367. [PubMed] [Google Scholar]
- 40.Roberts A, Buonocore L, Price R, Forman J, Rose JK. Attenuated vesicular stomatitis viruses as vaccine vectors. J Virol. 1999;73:3723–3732. doi: 10.1128/JVI.73.5.3723-3732.1999. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 41.Lichty BD, Power AT, Stojdl DF, Bell JC. Vesicular stomatitis virus: Re-inventing the bullet. Trends Mol Med. 2004;10:210–216. doi: 10.1016/j.molmed.2004.03.003. [DOI] [PubMed] [Google Scholar]
- 42.Hastie E, Cataldi M, Marriott I, Grdzelishvili VZ. Understanding and altering cell tropism of vesicular stomatitis virus. Virus Res. 2013;176:16–32. doi: 10.1016/j.virusres.2013.06.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 43.van den Pol AN, Ozduman K, Wollmann G, Ho WS, Simon I, Yao Y, et al. Viral strategies for studying the brain, including a replication-restricted self-amplifying delta-G vesicular stomatis virus that rapidly expresses transgenes in brain and can generate a multicolor Golgi-like expression. J Comp Neurol. 2009;516:456–481. doi: 10.1002/cne.22131. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 44.Beier KT, Saunders A, Oldenburg IA, Miyamichi K, Akhtar N, Luo L, et al. Anterograde or retrograde transsynaptic labeling of CNS neurons with vesicular stomatitis virus vectors. PNAS. 2011;108:15414–15419. doi: 10.1073/pnas.1110854108. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 45.Zingg B, Chou XL, Zhang ZG, Mesik L, Liang FX, Tao HW, et al. AAV-mediated anterograde transsynaptic tagging: Mapping corticocollicular input-defined neural pathways for defense behaviors. Neuron. 2017;93:33–47. doi: 10.1016/j.neuron.2016.11.045. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 46.Hollis Ii ER, Kadoya K, Hirsch M, Samulski RJ, Tuszynski MH. Efficient retrograde neuronal transduction utilizing self-complementary AAV1. Mol Ther. 2008;16:296–301. doi: 10.1038/sj.mt.6300367. [DOI] [PubMed] [Google Scholar]
- 47.Wickersham IR, Lyon DC, Barnard RJ, Mori T, Finke S, Conzelmann KK, et al. Monosynaptic restriction of transsynaptic tracing from single, genetically targeted neurons. Neuron. 2007;53:639–647. doi: 10.1016/j.neuron.2007.01.033. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 48.Wall NR, Wickersham IR, Cetin A, De La Parra M, Callaway EM. Monosynaptic circuit tracing in vivo through Cre-dependent targeting and complementation of modified rabies virus. Proc Natl Acad Sci USA. 2010;107:21848–21853. doi: 10.1073/pnas.1011756107. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 49.Wickersham IR, Finke S, Conzelmann KK, Callaway EM. Retrograde neuronal tracing with a deletion-mutant rabies virus. Nat Methods. 2007;4:47–49. doi: 10.1038/nmeth999. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 50.Schwarz LA, Miyamichi K, Gao XJ, Beier KT, Weissbourd B, DeLoach KE, et al. Viral-genetic tracing of the input-output organization of a central noradrenaline circuit. Nature. 2015;524:88–92. doi: 10.1038/nature14600. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 51.Chatterjee S, Sullivan HA, MacLennan BJ, Xu R, Hou YY, Lavin TK, et al. Nontoxic, double-deletion-mutant rabies viral vectors for retrograde targeting of projection neurons. Nat Neurosci. 2018;21:638–646. doi: 10.1038/s41593-018-0091-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 52.Albertini AA, Ruigrok RW, Blondel D. Rabies virus transcription and replication. Adv Virus Res. 2011;79:1–22. doi: 10.1016/B978-0-12-387040-7.00001-9. [DOI] [PubMed] [Google Scholar]
- 53.Boyden ES. A history of optogenetics: The development of tools for controlling brain circuits with light. F1000 Biol Rep. 2011;3:11. doi: 10.3410/B3-11. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 54.Deisseroth K. Optogenetics. Nat Methods. 2011;8:26–29. doi: 10.1038/nmeth.f.324. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 55.Matsuno-Yagi A, Mukohata Y. Two possible roles of bacteriorhodopsin; a comparative study of strains of Halobacterium halobium differing in pigmentation. Biochem Biophys Res Commun. 1977;78:237–243. doi: 10.1016/0006-291X(77)91245-1. [DOI] [PubMed] [Google Scholar]
- 56.Nagel G, Ollig D, Fuhrmann M, Kateriya S, Musti AM, Bamberg E, et al. Channelrhodopsin-1: a light-gated proton channel in green algae. Science. 2002;296:2395–2398. doi: 10.1126/science.1072068. [DOI] [PubMed] [Google Scholar]
- 57.Boyden ES, Zhang F, Bamberg E, Nagel G, Deisseroth K. Millisecond-timescale, genetically targeted optical control of neural activity. Nat Neurosci. 2005;8:1263–1268. doi: 10.1038/nn1525. [DOI] [PubMed] [Google Scholar]
- 58.Aravanis AM, Wang LP, Zhang F, Meltzer LA, Mogri MZ, Schneider MB, et al. An optical neural interface: In vivo control of rodent motor cortex with integrated fiberoptic and optogenetic technology. J Neural Eng. 2007;4:S143–S156. doi: 10.1088/1741-2560/4/3/S02. [DOI] [PubMed] [Google Scholar]
- 59.Adamantidis AR, Zhang F, Aravanis AM, Deisseroth K, de Lecea L. Neural substrates of awakening probed with optogenetic control of hypocretin neurons. Nature. 2007;450:420–424. doi: 10.1038/nature06310. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 60.Kato HE, Zhang F, Yizhar O, Ramakrishnan C, Nishizawa T, Hirata K, et al. Crystal structure of the channelrhodopsin light-gated cation channel. Nature. 2012;482:369–374. doi: 10.1038/nature10870. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 61.Deisseroth K. Optogenetics: 10 years of microbial opsins in neuroscience. Nat Neurosci. 2015;18:1213–1225. doi: 10.1038/nn.4091. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 62.Schneider F, Grimm C, Hegemann P. Biophysics of channelrhodopsin. Annu Rev Biophys. 2015;44:167–186. doi: 10.1146/annurev-biophys-060414-034014. [DOI] [PubMed] [Google Scholar]
- 63.Nagel G, Brauner M, Liewald JF, Adeishvili N, Bamberg E, Gottschalk A. Light activation of channelrhodopsin-2 in excitable cells of Caenorhabditis elegans triggers rapid behavioral responses. Curr Biol. 2005;15:2279–2284. doi: 10.1016/j.cub.2005.11.032. [DOI] [PubMed] [Google Scholar]
- 64.Lin JY, Lin MZ, Steinbach P, Tsien RY. Characterization of engineered channelrhodopsin variants with improved properties and kinetics. Biophys J. 2009;96:1803–1814. doi: 10.1016/j.bpj.2008.11.034. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 65.Gunaydin LA, Yizhar O, Berndt A, Sohal VS, Deisseroth K, Hegemann P. Ultrafast optogenetic control. Nat Neurosci. 2010;13:387–392. doi: 10.1038/nn.2495. [DOI] [PubMed] [Google Scholar]
- 66.Zhang F, Prigge M, Beyrière F, Tsunoda SP, Mattis J, Yizhar O, et al. Red-shifted optogenetic excitation: A tool for fast neural control derived from Volvox carteri. Nat Neurosci. 2008;11:631–633. doi: 10.1038/nn.2120. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 67.Lin JY, Knutsen PM, Muller A, Kleinfeld D, Tsien RY. ReaChR: a red-shifted variant of channelrhodopsin enables deep transcranial optogenetic excitation. Nat Neurosci. 2013;16:1499–1508. doi: 10.1038/nn.3502. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 68.Hochbaum DR, Zhao YX, Farhi SL, Klapoetke N, Werley CA, Kapoor V, et al. All-optical electrophysiology in mammalian neurons using engineered microbial rhodopsins. Nat Methods. 2014;11:825–833. doi: 10.1038/nmeth.3000. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 69.Chow BY, Han X, Dobry AS, Qian XF, Chuong AS, Li MJ, et al. High-performance genetically targetable optical neural silencing by light-driven proton pumps. Nature. 2010;463:98–102. doi: 10.1038/nature08652. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 70.Wietek J, Wiegert JS, Adeishvili N, Schneider F, Watanabe H, Tsunoda SP, et al. Conversion of channelrhodopsin into a light-gated chloride channel. Science. 2014;344:409–412. doi: 10.1126/science.1249375. [DOI] [PubMed] [Google Scholar]
- 71.Atasoy D, Sternson SM. Chemogenetic tools for causal cellular and neuronal biology. Physiol Rev. 2018;98:391–418. doi: 10.1152/physrev.00009.2017. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 72.Conklin BR, Hsiao EC, Claeysen S, Dumuis A, Srinivasan S, Forsayeth JR, et al. Engineering GPCR signaling pathways with RASSLs. Nat Methods. 2008;5:673–678. doi: 10.1038/nmeth.1232. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 73.Alexander GM, Rogan SC, Abbas AI, Armbruster BN, Pei Y, Allen JA, et al. Remote control of neuronal activity in transgenic mice expressing evolved G protein-coupled receptors. Neuron. 2009;63:27–39. doi: 10.1016/j.neuron.2009.06.014. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 74.Stachniak TJ, Ghosh A, Sternson SM. Chemogenetic synaptic silencing of neural circuits localizes a hypothalamus→midbrain pathway for feeding behavior. Neuron. 2014;82:797–808. doi: 10.1016/j.neuron.2014.04.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 75.Gomez JL, Bonaventura J, Lesniak W, Mathews WB, Sysa-Shah P, Rodriguez LA, et al. Chemogenetics revealed: DREADD occupancy and activation via converted clozapine. Science. 2017;357:503–507. doi: 10.1126/science.aan2475. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 76.Todd WD, Fenselau H, Wang JL, Zhang R, Machado NL, Venner A, et al. A hypothalamic circuit for the circadian control of aggression. Nat Neurosci. 2018;21:717–724. doi: 10.1038/s41593-018-0126-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 77.Rodriguez GA, Barrett GM, Duff KE, Hussaini SA. Chemogenetic attenuation of neuronal activity in the entorhinal cortex reduces Aβ and tau pathology in the Hippocampus. PLoS Biol. 2020;18:e3000851. doi: 10.1371/journal.pbio.3000851. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 78.Braak H, Braak E. Neuropathological stageing of Alzheimer-related changes. Acta Neuropathol. 1991;82:239–259. doi: 10.1007/BF00308809. [DOI] [PubMed] [Google Scholar]
- 79.Serrano-Pozo A, Frosch MP, Masliah E, Hyman BT. Neuropathological alterations in Alzheimer disease. Cold Spring Harb Perspect Med. 2011;1:a006189. doi: 10.1101/cshperspect.a006189. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 80.Devanand DP, Pradhaban G, Liu X, Khandji A, De Santi S, Segal S, et al. Hippocampal and entorhinal atrophy in mild cognitive impairment: Prediction of Alzheimer disease. Neurology. 2007;68:828–836. doi: 10.1212/01.wnl.0000256697.20968.d7. [DOI] [PubMed] [Google Scholar]
- 81.Busche MA, Hyman BT. Synergy between amyloid-β and tau in Alzheimer's disease. Nat Neurosci. 2020;23:1183–1193. doi: 10.1038/s41593-020-0687-6. [DOI] [PubMed] [Google Scholar]
- 82.Zhou YY, Song WM, Andhey PS, Swain A, Levy T, Miller KR, et al. Author Correction: Human and mouse single-nucleus transcriptomics reveal TREM2-dependent and TREM2-independent cellular responses in Alzheimer's disease. Nat Med. 2020;26:981. doi: 10.1038/s41591-020-0922-4. [DOI] [PubMed] [Google Scholar]
- 83.Huijbers W, Mormino EC, Schultz AP, Wigman S, Ward AM, Larvie M, et al. Amyloid-β deposition in mild cognitive impairment is associated with increased hippocampal activity, atrophy and clinical progression. Brain. 2015;138:1023–1035. doi: 10.1093/brain/awv007. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 84.Busche MA, Eichhoff G, Adelsberger H, Abramowski D, Wiederhold KH, Haass C, et al. Clusters of hyperactive neurons near amyloid plaques in a mouse model of Alzheimer's disease. Science. 2008;321:1686–1689. doi: 10.1126/science.1162844. [DOI] [PubMed] [Google Scholar]
- 85.Zott B, Simon MM, Hong W, Unger F, Chen-Engerer HJ, Frosch MP, et al. A vicious cycle of β amyloid-dependent neuronal hyperactivation. Science. 2019;365:559–565. doi: 10.1126/science.aay0198. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 86.Matos M, Augusto E, Machado NJ, dos Santos-Rodrigues A, Cunha RA, Agostinho P. Astrocytic adenosine A2A receptors control the amyloid-β peptide-induced decrease of glutamate uptake. J Alzheimers Dis. 2012;31:555–567. doi: 10.3233/JAD-2012-120469. [DOI] [PubMed] [Google Scholar]
- 87.Canas PM, Porciúncula LO, Cunha GM, Silva CG, Machado NJ, Oliveira JM, et al. Adenosine A2A receptor blockade prevents synaptotoxicity and memory dysfunction caused by beta-amyloid peptides via p38 mitogen-activated protein kinase pathway. J Neurosci. 2009;29:14741–14751. doi: 10.1523/JNEUROSCI.3728-09.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 88.Wang DY, Govindaiah G, Liu RJ, De Arcangelis V, Cox CL, Xiang YK. Binding of amyloid beta peptide to beta2 adrenergic receptor induces PKA-dependent AMPA receptor hyperactivity. FASEB J. 2010;24:3511–3521. doi: 10.1096/fj.10-156661. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 89.Huang Y, Shen W, Su J, Cheng B, Li D, Liu G, et al. Modulating the balance of synaptic and extrasynaptic NMDA receptors shows positive effects against amyloid-β-induced neurotoxicity. J Alzheimers Dis. 2017;57:885–897. doi: 10.3233/JAD-161186. [DOI] [PubMed] [Google Scholar]
- 90.Hardingham GE, Fukunaga Y, Bading H. Extrasynaptic NMDARs oppose synaptic NMDARs by triggering CREB shut-off and cell death pathways. Nat Neurosci. 2002;5:405–414. doi: 10.1038/nn835. [DOI] [PubMed] [Google Scholar]
- 91.Abramov E, Dolev I, Fogel H, Ciccotosto GD, Ruff E, Slutsky I. Amyloid-beta as a positive endogenous regulator of release probability at hippocampal synapses. Nat Neurosci. 2009;12:1567–1576. doi: 10.1038/nn.2433. [DOI] [PubMed] [Google Scholar]
- 92.He Y, Wei MD, Wu Y, Qin HP, Li WN, Ma XL, et al. Amyloid β oligomers suppress excitatory transmitter release via presynaptic depletion of phosphatidylinositol-4, 5-bisphosphate. Nat Commun. 2019;10:1193. doi: 10.1038/s41467-019-09114-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 93.Talantova M, Sanz-Blasco S, Zhang XF, Xia P, Akhtar MW, Okamoto S, et al. Aβ induces astrocytic glutamate release, extrasynaptic NMDA receptor activation, and synaptic loss. Proc Natl Acad Sci USA. 2013;110:E2518–E2527. doi: 10.1073/pnas.1306832110. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 94.Shankar GM, Li SM, Mehta TH, Garcia-Munoz A, Shepardson NE, Smith I, et al. Amyloid-beta protein dimers isolated directly from Alzheimer's brains impair synaptic plasticity and memory. Nat Med. 2008;14:837–842. doi: 10.1038/nm1782. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 95.Leshchyns'ka I, Liew HT, Shepherd C, Halliday GM, Stevens CH, Ke YD, et al. Aβ-dependent reduction of NCAM2-mediated synaptic adhesion contributes to synapse loss in Alzheimer's disease. Nat Commun. 2015;6:8836. doi: 10.1038/ncomms9836. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 96.Hsieh H, Boehm J, Sato C, Iwatsubo T, Tomita T, Sisodia S, et al. AMPAR removal underlies Abeta-induced synaptic depression and dendritic spine loss. Neuron. 2006;52:831–843. doi: 10.1016/j.neuron.2006.10.035. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 97.Hrynchak MV, Rierola M, Golovyashkina N, Penazzi L, Pump WC, David B, et al. Chronic presence of oligomeric aβ differentially modulates spine parameters in the Hippocampus and cortex of mice with low APP transgene expression. Front Synaptic Neurosci. 2020;12:16. doi: 10.3389/fnsyn.2020.00016. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 98.Ting JT, Kelley BG, Lambert TJ, Cook DG, Sullivan JM. Amyloid precursor protein overexpression depresses excitatory transmission through both presynaptic and postsynaptic mechanisms. Proc Natl Acad Sci USA. 2007;104:353–358. doi: 10.1073/pnas.0608807104. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 99.Lin L, Liu AY, Li HQ, Feng J, Yan Z. Inhibition of histone methyltransferases EHMT1/2 reverses amyloid-β-induced loss of AMPAR currents in human stem cell-derived cortical neurons. J Alzheimers Dis. 2019;70:1175–1185. doi: 10.3233/JAD-190190. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 100.Du YH, Fu M, Huang ZL, Tian X, Li JJ, Pang YY, et al. TRPV1 activation alleviates cognitive and synaptic plasticity impairments through inhibiting AMPAR endocytosis in APP23/PS45 mouse model of Alzheimer's disease. Aging Cell. 2020;19:e13113. doi: 10.1111/acel.13113. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 101.Rodrigues EM, Scudder SL, Goo MS, Patrick GN. Aβ-induced synaptic alterations require the E3 ubiquitin ligase Nedd4-1. J Neurosci. 2016;36:1590–1595. doi: 10.1523/JNEUROSCI.2964-15.2016. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 102.Snyder EM, Nong Y, Almeida CG, Paul S, Moran T, Choi EY, et al. Regulation of NMDA receptor trafficking by amyloid-beta. Nat Neurosci. 2005;8:1051–1058. doi: 10.1038/nn1503. [DOI] [PubMed] [Google Scholar]
- 103.Hsu WL, Ma YL, Hsieh DY, Liu YC, Lee EH. STAT1 negatively regulates spatial memory formation and mediates the memory-impairing effect of Aβ. Neuropsychopharmacology. 2014;39:746–758. doi: 10.1038/npp.2013.263. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 104.Yin YL, Gao D, Wang YL, Wang ZH, Wang X, Ye JW, et al. Tau accumulation induces synaptic impairment and memory deficit by calcineurin-mediated inactivation of nuclear CaMKIV/CREB signaling. Proc Natl Acad Sci USA. 2016;113:E3773–E3781. doi: 10.1073/pnas.1604519113. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 105.Li XG, Hong XY, Wang YL, Zhang SJ, Zhang JF, Li XC, et al. Tau accumulation triggers STAT1-dependent memory deficits by suppressing NMDA receptor expression. EMBO Rep. 2019;20:e47202. doi: 10.15252/embr.201847202. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 106.Ye JW, Yin YL, Liu HH, Fang L, Tao XQ, Wei LY, et al. Tau inhibits PKA by nuclear proteasome-dependent PKAR2α elevation with suppressed CREB/GluA1 phosphorylation. Aging Cell. 2020;19:e13055. doi: 10.1111/acel.13055. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 107.Ittner LM, Ke YD, Delerue F, Bi M, Gladbach A, van Eersel J, et al. Dendritic function of tau mediates amyloid-beta toxicity in Alzheimer's disease mouse models. Cell. 2010;142:387–397. doi: 10.1016/j.cell.2010.06.036. [DOI] [PubMed] [Google Scholar]
- 108.Luo HB, Xia YY, Shu XJ, Liu ZC, Feng Y, Liu XH, et al. SUMOylation at K340 inhibits tau degradation through deregulating its phosphorylation and ubiquitination. Proc Natl Acad Sci USA. 2014;111:16586–16591. doi: 10.1073/pnas.1417548111. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 109.Feng Q, Luo Y, Zhang XN, Yang XF, Hong XY, Sun DS, et al. MAPT/Tau accumulation represses autophagy flux by disrupting IST1-regulated ESCRT-III complex formation: A vicious cycle in Alzheimer neurodegeneration. Autophagy. 2020;16:641–658. doi: 10.1080/15548627.2019.1633862. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 110.Wang X, Liu EJ, Liu Q, Li SH, Li T, Zhou QZ, et al. Tau acetylation in entorhinal cortex induces its chronic hippocampal propagation and cognitive deficits in mice. J Alzheimers Dis. 2020;77:241–255. doi: 10.3233/JAD-200529. [DOI] [PubMed] [Google Scholar]
- 111.Ye JW, Yin Y, Yin YL, Zhang HQ, Wan HL, Wang L, et al. Tau-induced upregulation of C/EBPβ-TRPC1-SOCE signaling aggravates tauopathies: A vicious cycle in Alzheimer neurodegeneration. Aging Cell. 2020;19:e13209. doi: 10.1111/acel.13209. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 112.Hoover BR, Reed MN, Su JJ, Penrod RD, Kotilinek LA, Grant MK, et al. Tau mislocalization to dendritic spines mediates synaptic dysfunction independently of neurodegeneration. Neuron. 2010;68:1067–1081. doi: 10.1016/j.neuron.2010.11.030. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 113.Tracy TE, Sohn PD, Minami SS, Wang C, Min SW, Li YQ, et al. Acetylated tau obstructs KIBRA-mediated signaling in synaptic plasticity and promotes tauopathy-related memory loss. Neuron. 2016;90:245–260. doi: 10.1016/j.neuron.2016.03.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 114.Zhao XH, Kotilinek LA, Smith B, Hlynialuk C, Zahs K, Ramsden M, et al. Caspase-2 cleavage of tau reversibly impairs memory. Nat Med. 2016;22:1268–1276. doi: 10.1038/nm.4199. [DOI] [PubMed] [Google Scholar]
- 115.Hatch RJ, Wei Y, Xia D, Götz J. Hyperphosphorylated tau causes reduced hippocampal CA1 excitability by relocating the axon initial segment. Acta Neuropathol. 2017;133:717–730. doi: 10.1007/s00401-017-1674-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 116.Sohn PD, Huang CT, Yan R, Fan L, Tracy TE, Camargo CM, et al. Pathogenic tau impairs axon initial segment plasticity and excitability homeostasis. Neuron. 2019;104:458–470.e5. doi: 10.1016/j.neuron.2019.08.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 117.Chen TW, Wardill TJ, Sun Y, Pulver SR, Renninger SL, Baohan A, et al. Ultrasensitive fluorescent proteins for imaging neuronal activity. Nature. 2013;499:295–300. doi: 10.1038/nature12354. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 118.Jimenez JC, Su K, Goldberg AR, Luna VM, Biane JS, Ordek G, et al. Anxiety cells in a hippocampal-hypothalamic circuit. Neuron. 2018;97:670–683.e6. doi: 10.1016/j.neuron.2018.01.016. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 119.Granholm E, Butters N. Associative encoding and retrieval in Alzheimer's and Huntington's disease. Brain Cogn. 1988;7:335–347. doi: 10.1016/0278-2626(88)90007-3. [DOI] [PubMed] [Google Scholar]
- 120.Roy DS, Arons A, Mitchell TI, Pignatelli M, Ryan TJ, Tonegawa S. Memory retrieval by activating engram cells in mouse models of early Alzheimer's disease. Nature. 2016;531:508–512. doi: 10.1038/nature17172. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 121.He QY, Wang JH, Hu HL. Illuminating the activated brain: Emerging activity-dependent tools to capture and control functional neural circuits. Neurosci Bull. 2019;35:369–377. doi: 10.1007/s12264-018-0291-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 122.Yang Y, Wang ZH, Jin S, Gao D, Liu N, Chen SP, et al. Opposite monosynaptic scaling of BLP-vCA1 inputs governs hopefulness- and helplessness-modulated spatial learning and memory. Nat Commun. 2016;7:11935. doi: 10.1038/ncomms11935. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 123.Wu JW, Hussaini SA, Bastille IM, Rodriguez GA, Mrejeru A, Rilett K, et al. Neuronal activity enhances tau propagation and tau pathology in vivo. Nat Neurosci. 2016;19:1085–1092. doi: 10.1038/nn.4328. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 124.Garcia-Marin V, Blazquez-Llorca L, Rodriguez JR, Boluda S, Muntane G, Ferrer I, et al. Diminished perisomatic GABAergic terminals on cortical neurons adjacent to amyloid plaques. Front Neuroanat. 2009;3:28. doi: 10.3389/neuro.05.028.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 125.Ramos-Miguel A, Hercher C, Beasley CL, Barr AM, Bayer TA, Falkai P, et al. Loss of Munc18-1 long splice variant in GABAergic terminals is associated with cognitive decline and increased risk of dementia in a community sample. Mol Neurodegener. 2015;10:65. doi: 10.1186/s13024-015-0061-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 126.Bell KF, de Kort GJ, Steggerda S, Shigemoto R, Ribeiro-da-Silva A, Cuello AC. Structural involvement of the glutamatergic presynaptic boutons in a transgenic mouse model expressing early onset amyloid pathology. Neurosci Lett. 2003;353:143–147. doi: 10.1016/j.neulet.2003.09.027. [DOI] [PubMed] [Google Scholar]
- 127.Bell KF, Ducatenzeiler A, Ribeiro-da-Silva A, Duff K, Bennett DA, Cuello AC. The amyloid pathology progresses in a neurotransmitter-specific manner. Neurobiol Aging. 2006;27:1644–1657. doi: 10.1016/j.neurobiolaging.2005.09.034. [DOI] [PubMed] [Google Scholar]
- 128.Sanchez-Mejias E, Nuñez-Diaz C, Sanchez-Varo R, Gomez-Arboledas A, Garcia-Leon JA, Fernandez-Valenzuela JJ, et al. Distinct disease-sensitive GABAergic neurons in the perirhinal cortex of Alzheimer's mice and patients. Brain Pathol. 2020;30:345–363. doi: 10.1111/bpa.12785. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 129.Ramos B, Baglietto-Vargas D, del Rio JC, Moreno-Gonzalez I, Santa-Maria C, Jimenez S, et al. Early neuropathology of somatostatin/NPY GABAergic cells in the Hippocampus of a PS1xAPP transgenic model of Alzheimer's disease. Neurobiol Aging. 2006;27:1658–1672. doi: 10.1016/j.neurobiolaging.2005.09.022. [DOI] [PubMed] [Google Scholar]
- 130.Zallo F, Gardenal E, Verkhratsky A, Rodríguez JJ. Loss of calretinin and parvalbumin positive interneurones in the hippocampal CA1 of aged Alzheimer's disease mice. Neurosci Lett. 2018;681:19–25. doi: 10.1016/j.neulet.2018.05.027. [DOI] [PubMed] [Google Scholar]
- 131.Nykänen NP, Kysenius K, Sakha P, Tammela P, Huttunen HJ. Γ-Aminobutyric acid type A (GABAA) receptor activation modulates tau phosphorylation. J Biol Chem. 2012;287:6743–6752. doi: 10.1074/jbc.M111.309385. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 132.Nilsen LH, Rae C, Ittner LM, Götz J, Sonnewald U. Glutamate metabolism is impaired in transgenic mice with tau hyperphosphorylation. J Cereb Blood Flow Metab. 2013;33:684–691. doi: 10.1038/jcbfm.2012.212. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 133.Hazra A, Gu F, Aulakh A, Berridge C, Eriksen JL, Ziburkus J. Inhibitory neuron and hippocampal circuit dysfunction in an aged mouse model of Alzheimer's disease. PLoS ONE. 2013;8:e64318. doi: 10.1371/journal.pone.0064318. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 134.Zheng J, Li HL, Tian N, Liu F, Wang L, Yin YL, et al. Interneuron accumulation of phosphorylated tau impairs adult hippocampal neurogenesis by suppressing GABAergic transmission. Cell Stem Cell. 2020;26:462–466. doi: 10.1016/j.stem.2020.01.021. [DOI] [PubMed] [Google Scholar]
- 135.Yang X, Yao C, Tian T, Li X, Yan H, Wu J, et al. A novel mechanism of memory loss in Alzheimer's disease mice via the degeneration of entorhinal-CA1 synapses. Mol Psychiatry. 2018;23:199–210. doi: 10.1038/mp.2016.151. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 136.Bienvenu TC, Busti D, Magill PJ, Ferraguti F, Capogna M. Cell-type-specific recruitment of amygdala interneurons to hippocampal Theta rhythm and noxious stimuli in vivo. Neuron. 2012;74:1059–1074. doi: 10.1016/j.neuron.2012.04.022. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 137.Loreth D, Ozmen L, Revel FG, Knoflach F, Wetzel P, Frotscher M, et al. Selective degeneration of septal and hippocampal GABAergic neurons in a mouse model of amyloidosis and tauopathy. Neurobiol Dis. 2012;47:1–12. doi: 10.1016/j.nbd.2012.03.011. [DOI] [PubMed] [Google Scholar]
- 138.Wang Y, Romani S, Lustig B, Leonardo A, Pastalkova E. Theta sequences are essential for internally generated hippocampal firing fields. Nat Neurosci. 2015;18:282–288. doi: 10.1038/nn.3904. [DOI] [PubMed] [Google Scholar]
- 139.Wikenheiser AM, Redish AD. Hippocampal Theta sequences reflect current goals. Nat Neurosci. 2015;18:289–294. doi: 10.1038/nn.3909. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 140.Cheng JH, Ji DY. Rigid firing sequences undermine spatial memory codes in a neurodegenerative mouse model. Elife. 2013;2:e00647. doi: 10.7554/eLife.00647. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 141.Villette V, Poindessous-Jazat F, Simon A, Léna C, Roullot E, Bellessort B, et al. Decreased rhythmic GABAergic septal activity and memory-associated Theta oscillations after hippocampal amyloid-beta pathology in the rat. J Neurosci. 2010;30:10991–11003. doi: 10.1523/JNEUROSCI.6284-09.2010. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 142.Martorell AJ, Paulson AL, Suk HJ, Abdurrob F, Drummond GT, Guan W, et al. Multi-sensory gamma stimulation ameliorates Alzheimer's-associated pathology and improves cognition. Cell. 2019;177:256–271.e22. doi: 10.1016/j.cell.2019.02.014. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 143.Jun H, Bramian A, Soma S, Saito T, Saido TC, Igarashi KM. Disrupted place cell remapping and impaired grid cells in a knockin model of Alzheimer's disease. Neuron. 2020;107:1095–1112.e6. doi: 10.1016/j.neuron.2020.06.023. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 144.Herrmann CS, Demiralp T. Human EEG gamma oscillations in neuropsychiatric disorders. Clin Neurophysiol. 2005;116:2719–2733. doi: 10.1016/j.clinph.2005.07.007. [DOI] [PubMed] [Google Scholar]
- 145.Yamamoto J, Suh J, Takeuchi D, Tonegawa S. Successful execution of working memory linked to synchronized high-frequency gamma oscillations. Cell. 2014;157:845–857. doi: 10.1016/j.cell.2014.04.009. [DOI] [PubMed] [Google Scholar]
- 146.Mikulovic S, Restrepo CE, Siwani S, Bauer P, Pupe S, Tort ABL, et al. Ventral hippocampal OLM cells control type 2 Theta oscillations and response to predator odor. Nat Commun. 2018;9:3638. doi: 10.1038/s41467-018-05907-w. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 147.Iaccarino HF, Singer AC, Martorell AJ, Rudenko A, Gao F, Gillingham TZ, et al. Author Correction: Gamma frequency entrainment attenuates amyloid load and modifies microglia. Nature. 2018;562:E1. doi: 10.1038/s41586-018-0351-4. [DOI] [PubMed] [Google Scholar]
- 148.Huh CY, Amilhon B, Ferguson KA, Manseau F, Torres-Platas SG, Peach JP, et al. Excitatory inputs determine phase-locking strength and spike-timing of CA1 stratum Oriens/Alveus parvalbumin and somatostatin interneurons during intrinsically generated hippocampal Theta rhythm. J Neurosci. 2016;36:6605–6622. doi: 10.1523/JNEUROSCI.3951-13.2016. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 149.Etter G, van der Veldt S, Manseau F, Zarrinkoub I, Trillaud-Doppia E, Williams S. Optogenetic gamma stimulation rescues memory impairments in an Alzheimer's disease mouse model. Nat Commun. 2019;10:5322. doi: 10.1038/s41467-019-13260-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 150.Chung H, Park K, Jang HJ, Kohl MM, Kwag J. Dissociation of somatostatin and parvalbumin interneurons circuit dysfunctions underlying hippocampal Theta and gamma oscillations impaired by amyloid β oligomers in vivo. Brain Struct Funct. 2020;225:935–954. doi: 10.1007/s00429-020-02044-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 151.Mesulam MM. The cholinergic innervation of the human cerebral cortex. Prog Brain Res. 2004;145:67–78. doi: 10.1016/S0079-6123(03)45004-8. [DOI] [PubMed] [Google Scholar]
- 152.Wilcock GK, Esiri MM, Bowen DM, Smith CC. Alzheimer's disease. Correlation of cortical choline acetyltransferase activity with the severity of dementia and histological abnormalities. J Neurol Sci. 1990;57:407–417. doi: 10.1016/0022-510X(82)90045-4. [DOI] [PubMed] [Google Scholar]
- 153.Kooi EJ, Prins M, Bajic N, Beliën JA, Gerritsen WH, van Horssen J, et al. Cholinergic imbalance in the multiple sclerosis Hippocampus. Acta Neuropathol. 2011;122:313–322. doi: 10.1007/s00401-011-0849-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 154.Yoon SY, Choi JU, Cho MH, Yang KM, Ha H, Chung IJ, et al. Α-secretase cleaved amyloid precursor protein (APP) accumulates in cholinergic dystrophic neurites in normal, aged Hippocampus. Neuropathol Appl Neurobiol. 2013;39:800–816. doi: 10.1111/nan.12032. [DOI] [PubMed] [Google Scholar]
- 155.Fang LQ, Duan JJ, Ran DZ, Fan ZH, Yan Y, Huang NY, et al. Amyloid-β depresses excitatory cholinergic synaptic transmission in Drosophila. Neurosci Bull. 2012;28:585–594. doi: 10.1007/s12264-012-1267-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 156.Berger-Sweeney J, Stearns NA, Murg SL, Floerke-Nashner LR, Lappi DA, Baxter MG. Selective immunolesions of cholinergic neurons in mice: Effects on neuroanatomy, neurochemistry, and behavior. J Neurosci. 2001;21:8164–8173. doi: 10.1523/JNEUROSCI.21-20-08164.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 157.Lai MK, Tsang SW, Alder JT, Keene J, Hope T, Esiri MM, et al. Loss of serotonin 5-HT2A receptors in the postmortem temporal cortex correlates with rate of cognitive decline in Alzheimer's disease. Psychopharmacology (Berl) 2005;179:673–677. doi: 10.1007/s00213-004-2077-2. [DOI] [PubMed] [Google Scholar]
- 158.Lai MK, Tsang SW, Francis PT, Keene J, Hope T, Esiri MM, et al. Postmortem serotoninergic correlates of cognitive decline in Alzheimer's disease. Neuroreport. 2002;13:1175–1178. doi: 10.1097/00001756-200207020-00021. [DOI] [PubMed] [Google Scholar]
- 159.Yamamoto T, Hirano A. Nucleus raphe dorsalis in Alzheimer's disease: Neurofibrillary tangles and loss of large neurons. Ann Neurol. 1985;17:573–577. doi: 10.1002/ana.410170608. [DOI] [PubMed] [Google Scholar]
- 160.Mowla A, Mosavinasab M, Haghshenas H, Borhani Haghighi A. Does serotonin augmentation have any effect on cognition and activities of daily living in Alzheimer's dementia? A double-blind, placebo-controlled clinical trial. J Clin Psychopharmacol. 2007;27:484–487. doi: 10.1097/jcp.0b013e31814b98c1. [DOI] [PubMed] [Google Scholar]
- 161.Mokler DJ, Lariviere D, Johnson DW, Theriault NL, Bronzino JD, Dixon M, et al. Serotonin neuronal release from dorsal Hippocampus following electrical stimulation of the dorsal and Median raphé nuclei in conscious rats. Hippocampus. 1998;8:262–273. doi: 10.1002/(SICI)1098-1063(1998)8:3<262::AID-HIPO8>3.0.CO;2-L. [DOI] [PubMed] [Google Scholar]
- 162.Yoshida K, Drew MR, Mimura M, Tanaka KF. Serotonin-mediated inhibition of ventral Hippocampus is required for sustained goal-directed behavior. Nat Neurosci. 2019;22:770–777. doi: 10.1038/s41593-019-0376-5. [DOI] [PubMed] [Google Scholar]
- 163.Abela AR, Browne CJ, Sargin D, Prevot TD, Ji XD, Li Z, et al. Median raphe serotonin neurons promote anxiety-like behavior via inputs to the dorsal Hippocampus. Neuropharmacology. 2020;168:107985. doi: 10.1016/j.neuropharm.2020.107985. [DOI] [PubMed] [Google Scholar]
- 164.Teixeira CM, Rosen ZB, Suri D, Sun Q, Hersh M, Sargin D, et al. Hippocampal 5-HT input regulates memory formation and schaffer collateral excitation. Neuron. 2018;98:992–1004.e4. doi: 10.1016/j.neuron.2018.04.030. [DOI] [PMC free article] [PubMed] [Google Scholar]